ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/tengDissertation/Introduction.tex
Revision: 2872
Committed: Wed Jun 21 16:28:25 2006 UTC (18 years ago) by tim
Content type: application/x-tex
File size: 77088 byte(s)
Log Message:
correction to Chapter one

File Contents

# User Rev Content
1 tim 2685 \chapter{\label{chapt:introduction}INTRODUCTION AND THEORETICAL BACKGROUND}
2    
3 tim 2693 \section{\label{introSection:classicalMechanics}Classical
4     Mechanics}
5 tim 2685
6 tim 2692 Closely related to Classical Mechanics, Molecular Dynamics
7     simulations are carried out by integrating the equations of motion
8     for a given system of particles. There are three fundamental ideas
9 tim 2819 behind classical mechanics. Firstly, one can determine the state of
10 tim 2692 a mechanical system at any time of interest; Secondly, all the
11     mechanical properties of the system at that time can be determined
12     by combining the knowledge of the properties of the system with the
13     specification of this state; Finally, the specification of the state
14     when further combine with the laws of mechanics will also be
15     sufficient to predict the future behavior of the system.
16 tim 2685
17 tim 2693 \subsection{\label{introSection:newtonian}Newtonian Mechanics}
18 tim 2694 The discovery of Newton's three laws of mechanics which govern the
19     motion of particles is the foundation of the classical mechanics.
20 tim 2819 Newton's first law defines a class of inertial frames. Inertial
21 tim 2694 frames are reference frames where a particle not interacting with
22     other bodies will move with constant speed in the same direction.
23 tim 2819 With respect to inertial frames, Newton's second law has the form
24 tim 2694 \begin{equation}
25 tim 2819 F = \frac {dp}{dt} = \frac {mdv}{dt}
26 tim 2694 \label{introEquation:newtonSecondLaw}
27     \end{equation}
28     A point mass interacting with other bodies moves with the
29     acceleration along the direction of the force acting on it. Let
30 tim 2702 $F_{ij}$ be the force that particle $i$ exerts on particle $j$, and
31     $F_{ji}$ be the force that particle $j$ exerts on particle $i$.
32 tim 2819 Newton's third law states that
33 tim 2694 \begin{equation}
34 tim 2702 F_{ij} = -F_{ji}
35 tim 2694 \label{introEquation:newtonThirdLaw}
36     \end{equation}
37 tim 2692
38 tim 2694 Conservation laws of Newtonian Mechanics play very important roles
39     in solving mechanics problems. The linear momentum of a particle is
40     conserved if it is free or it experiences no force. The second
41     conservation theorem concerns the angular momentum of a particle.
42     The angular momentum $L$ of a particle with respect to an origin
43     from which $r$ is measured is defined to be
44     \begin{equation}
45     L \equiv r \times p \label{introEquation:angularMomentumDefinition}
46     \end{equation}
47     The torque $\tau$ with respect to the same origin is defined to be
48     \begin{equation}
49 tim 2819 \tau \equiv r \times F \label{introEquation:torqueDefinition}
50 tim 2694 \end{equation}
51     Differentiating Eq.~\ref{introEquation:angularMomentumDefinition},
52     \[
53     \dot L = \frac{d}{{dt}}(r \times p) = (\dot r \times p) + (r \times
54     \dot p)
55     \]
56     since
57     \[
58     \dot r \times p = \dot r \times mv = m\dot r \times \dot r \equiv 0
59     \]
60     thus,
61     \begin{equation}
62 tim 2819 \dot L = r \times \dot p = \tau
63 tim 2694 \end{equation}
64     If there are no external torques acting on a body, the angular
65     momentum of it is conserved. The last conservation theorem state
66 tim 2696 that if all forces are conservative, Energy
67     \begin{equation}E = T + V \label{introEquation:energyConservation}
68     \end{equation}
69     is conserved. All of these conserved quantities are
70     important factors to determine the quality of numerical integration
71 tim 2819 schemes for rigid bodies \cite{Dullweber1997}.
72 tim 2694
73 tim 2693 \subsection{\label{introSection:lagrangian}Lagrangian Mechanics}
74 tim 2692
75 tim 2819 Newtonian Mechanics suffers from two important limitations: motions
76     can only be described in cartesian coordinate systems. Moreover, It
77     become impossible to predict analytically the properties of the
78     system even if we know all of the details of the interaction. In
79     order to overcome some of the practical difficulties which arise in
80     attempts to apply Newton's equation to complex system, approximate
81     numerical procedures may be developed.
82 tim 2692
83 tim 2819 \subsubsection{\label{introSection:halmiltonPrinciple}\textbf{Hamilton's
84     Principle}}
85 tim 2692
86     Hamilton introduced the dynamical principle upon which it is
87 tim 2819 possible to base all of mechanics and most of classical physics.
88     Hamilton's Principle may be stated as follows,
89 tim 2692
90     The actual trajectory, along which a dynamical system may move from
91     one point to another within a specified time, is derived by finding
92     the path which minimizes the time integral of the difference between
93 tim 2819 the kinetic, $K$, and potential energies, $U$.
94 tim 2692 \begin{equation}
95     \delta \int_{t_1 }^{t_2 } {(K - U)dt = 0} ,
96 tim 2693 \label{introEquation:halmitonianPrinciple1}
97 tim 2692 \end{equation}
98    
99     For simple mechanical systems, where the forces acting on the
100 tim 2819 different parts are derivable from a potential, the Lagrangian
101     function $L$ can be defined as the difference between the kinetic
102     energy of the system and its potential energy,
103 tim 2692 \begin{equation}
104     L \equiv K - U = L(q_i ,\dot q_i ) ,
105     \label{introEquation:lagrangianDef}
106     \end{equation}
107     then Eq.~\ref{introEquation:halmitonianPrinciple1} becomes
108     \begin{equation}
109 tim 2693 \delta \int_{t_1 }^{t_2 } {L dt = 0} ,
110     \label{introEquation:halmitonianPrinciple2}
111 tim 2692 \end{equation}
112    
113 tim 2819 \subsubsection{\label{introSection:equationOfMotionLagrangian}\textbf{The
114     Equations of Motion in Lagrangian Mechanics}}
115 tim 2692
116 tim 2850 For a system of $f$ degrees of freedom, the equations of motion in
117     the Lagrangian form is
118 tim 2692 \begin{equation}
119     \frac{d}{{dt}}\frac{{\partial L}}{{\partial \dot q_i }} -
120     \frac{{\partial L}}{{\partial q_i }} = 0,{\rm{ }}i = 1, \ldots,f
121 tim 2693 \label{introEquation:eqMotionLagrangian}
122 tim 2692 \end{equation}
123     where $q_{i}$ is generalized coordinate and $\dot{q_{i}}$ is
124     generalized velocity.
125    
126 tim 2693 \subsection{\label{introSection:hamiltonian}Hamiltonian Mechanics}
127 tim 2692
128     Arising from Lagrangian Mechanics, Hamiltonian Mechanics was
129     introduced by William Rowan Hamilton in 1833 as a re-formulation of
130     classical mechanics. If the potential energy of a system is
131 tim 2819 independent of velocities, the momenta can be defined as
132 tim 2692 \begin{equation}
133     p_i = \frac{\partial L}{\partial \dot q_i}
134     \label{introEquation:generalizedMomenta}
135     \end{equation}
136 tim 2693 The Lagrange equations of motion are then expressed by
137 tim 2692 \begin{equation}
138 tim 2693 p_i = \frac{{\partial L}}{{\partial q_i }}
139     \label{introEquation:generalizedMomentaDot}
140     \end{equation}
141    
142     With the help of the generalized momenta, we may now define a new
143     quantity $H$ by the equation
144     \begin{equation}
145     H = \sum\limits_k {p_k \dot q_k } - L ,
146 tim 2692 \label{introEquation:hamiltonianDefByLagrangian}
147     \end{equation}
148     where $ \dot q_1 \ldots \dot q_f $ are generalized velocities and
149     $L$ is the Lagrangian function for the system.
150    
151 tim 2693 Differentiating Eq.~\ref{introEquation:hamiltonianDefByLagrangian},
152     one can obtain
153     \begin{equation}
154     dH = \sum\limits_k {\left( {p_k d\dot q_k + \dot q_k dp_k -
155     \frac{{\partial L}}{{\partial q_k }}dq_k - \frac{{\partial
156     L}}{{\partial \dot q_k }}d\dot q_k } \right)} - \frac{{\partial
157     L}}{{\partial t}}dt \label{introEquation:diffHamiltonian1}
158     \end{equation}
159     Making use of Eq.~\ref{introEquation:generalizedMomenta}, the
160     second and fourth terms in the parentheses cancel. Therefore,
161     Eq.~\ref{introEquation:diffHamiltonian1} can be rewritten as
162     \begin{equation}
163     dH = \sum\limits_k {\left( {\dot q_k dp_k - \dot p_k dq_k }
164     \right)} - \frac{{\partial L}}{{\partial t}}dt
165     \label{introEquation:diffHamiltonian2}
166     \end{equation}
167     By identifying the coefficients of $dq_k$, $dp_k$ and dt, we can
168     find
169     \begin{equation}
170 tim 2819 \frac{{\partial H}}{{\partial p_k }} = \dot {q_k}
171 tim 2693 \label{introEquation:motionHamiltonianCoordinate}
172     \end{equation}
173     \begin{equation}
174 tim 2819 \frac{{\partial H}}{{\partial q_k }} = - \dot {p_k}
175 tim 2693 \label{introEquation:motionHamiltonianMomentum}
176     \end{equation}
177     and
178     \begin{equation}
179     \frac{{\partial H}}{{\partial t}} = - \frac{{\partial L}}{{\partial
180     t}}
181     \label{introEquation:motionHamiltonianTime}
182     \end{equation}
183    
184     Eq.~\ref{introEquation:motionHamiltonianCoordinate} and
185     Eq.~\ref{introEquation:motionHamiltonianMomentum} are Hamilton's
186     equation of motion. Due to their symmetrical formula, they are also
187 tim 2786 known as the canonical equations of motions \cite{Goldstein2001}.
188 tim 2693
189 tim 2692 An important difference between Lagrangian approach and the
190     Hamiltonian approach is that the Lagrangian is considered to be a
191 tim 2819 function of the generalized velocities $\dot q_i$ and coordinates
192     $q_i$, while the Hamiltonian is considered to be a function of the
193     generalized momenta $p_i$ and the conjugate coordinates $q_i$.
194     Hamiltonian Mechanics is more appropriate for application to
195     statistical mechanics and quantum mechanics, since it treats the
196     coordinate and its time derivative as independent variables and it
197     only works with 1st-order differential equations\cite{Marion1990}.
198 tim 2692
199 tim 2696 In Newtonian Mechanics, a system described by conservative forces
200     conserves the total energy \ref{introEquation:energyConservation}.
201     It follows that Hamilton's equations of motion conserve the total
202     Hamiltonian.
203     \begin{equation}
204     \frac{{dH}}{{dt}} = \sum\limits_i {\left( {\frac{{\partial
205     H}}{{\partial q_i }}\dot q_i + \frac{{\partial H}}{{\partial p_i
206     }}\dot p_i } \right)} = \sum\limits_i {\left( {\frac{{\partial
207     H}}{{\partial q_i }}\frac{{\partial H}}{{\partial p_i }} -
208     \frac{{\partial H}}{{\partial p_i }}\frac{{\partial H}}{{\partial
209 tim 2698 q_i }}} \right) = 0} \label{introEquation:conserveHalmitonian}
210 tim 2696 \end{equation}
211    
212 tim 2693 \section{\label{introSection:statisticalMechanics}Statistical
213     Mechanics}
214 tim 2692
215 tim 2694 The thermodynamic behaviors and properties of Molecular Dynamics
216 tim 2692 simulation are governed by the principle of Statistical Mechanics.
217     The following section will give a brief introduction to some of the
218 tim 2700 Statistical Mechanics concepts and theorem presented in this
219     dissertation.
220 tim 2692
221 tim 2700 \subsection{\label{introSection:ensemble}Phase Space and Ensemble}
222 tim 2692
223 tim 2700 Mathematically, phase space is the space which represents all
224     possible states. Each possible state of the system corresponds to
225     one unique point in the phase space. For mechanical systems, the
226     phase space usually consists of all possible values of position and
227 tim 2819 momentum variables. Consider a dynamic system of $f$ particles in a
228     cartesian space, where each of the $6f$ coordinates and momenta is
229     assigned to one of $6f$ mutually orthogonal axes, the phase space of
230     this system is a $6f$ dimensional space. A point, $x = (q_1 , \ldots
231     ,q_f ,p_1 , \ldots ,p_f )$, with a unique set of values of $6f$
232     coordinates and momenta is a phase space vector.
233 tim 2700
234 tim 2850 %%%fix me
235 tim 2700 A microscopic state or microstate of a classical system is
236     specification of the complete phase space vector of a system at any
237     instant in time. An ensemble is defined as a collection of systems
238     sharing one or more macroscopic characteristics but each being in a
239     unique microstate. The complete ensemble is specified by giving all
240     systems or microstates consistent with the common macroscopic
241     characteristics of the ensemble. Although the state of each
242     individual system in the ensemble could be precisely described at
243     any instance in time by a suitable phase space vector, when using
244     ensembles for statistical purposes, there is no need to maintain
245     distinctions between individual systems, since the numbers of
246     systems at any time in the different states which correspond to
247     different regions of the phase space are more interesting. Moreover,
248     in the point of view of statistical mechanics, one would prefer to
249     use ensembles containing a large enough population of separate
250     members so that the numbers of systems in such different states can
251     be regarded as changing continuously as we traverse different
252     regions of the phase space. The condition of an ensemble at any time
253     can be regarded as appropriately specified by the density $\rho$
254     with which representative points are distributed over the phase
255 tim 2819 space. The density distribution for an ensemble with $f$ degrees of
256     freedom is defined as,
257 tim 2700 \begin{equation}
258     \rho = \rho (q_1 , \ldots ,q_f ,p_1 , \ldots ,p_f ,t).
259     \label{introEquation:densityDistribution}
260     \end{equation}
261     Governed by the principles of mechanics, the phase points change
262 tim 2819 their locations which would change the density at any time at phase
263     space. Hence, the density distribution is also to be taken as a
264 tim 2700 function of the time.
265    
266     The number of systems $\delta N$ at time $t$ can be determined by,
267     \begin{equation}
268     \delta N = \rho (q,p,t)dq_1 \ldots dq_f dp_1 \ldots dp_f.
269     \label{introEquation:deltaN}
270     \end{equation}
271 tim 2819 Assuming a large enough population of systems, we can sufficiently
272     approximate $\delta N$ without introducing discontinuity when we go
273     from one region in the phase space to another. By integrating over
274     the whole phase space,
275 tim 2700 \begin{equation}
276     N = \int { \ldots \int {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f
277     \label{introEquation:totalNumberSystem}
278     \end{equation}
279     gives us an expression for the total number of the systems. Hence,
280     the probability per unit in the phase space can be obtained by,
281     \begin{equation}
282     \frac{{\rho (q,p,t)}}{N} = \frac{{\rho (q,p,t)}}{{\int { \ldots \int
283     {\rho (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}.
284     \label{introEquation:unitProbability}
285     \end{equation}
286 tim 2850 With the help of Eq.~\ref{introEquation:unitProbability} and the
287     knowledge of the system, it is possible to calculate the average
288 tim 2700 value of any desired quantity which depends on the coordinates and
289     momenta of the system. Even when the dynamics of the real system is
290     complex, or stochastic, or even discontinuous, the average
291 tim 2819 properties of the ensemble of possibilities as a whole remaining
292     well defined. For a classical system in thermal equilibrium with its
293     environment, the ensemble average of a mechanical quantity, $\langle
294     A(q , p) \rangle_t$, takes the form of an integral over the phase
295     space of the system,
296 tim 2700 \begin{equation}
297     \langle A(q , p) \rangle_t = \frac{{\int { \ldots \int {A(q,p)\rho
298     (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}{{\int { \ldots \int {\rho
299     (q,p,t)dq_1 } ...dq_f dp_1 } ...dp_f }}
300     \label{introEquation:ensembelAverage}
301     \end{equation}
302    
303     There are several different types of ensembles with different
304     statistical characteristics. As a function of macroscopic
305 tim 2819 parameters, such as temperature \textit{etc}, the partition function
306     can be used to describe the statistical properties of a system in
307 tim 2700 thermodynamic equilibrium.
308    
309     As an ensemble of systems, each of which is known to be thermally
310 tim 2850 isolated and conserve energy, the Microcanonical ensemble (NVE) has
311     a partition function like,
312 tim 2700 \begin{equation}
313 tim 2706 \Omega (N,V,E) = e^{\beta TS} \label{introEquation:NVEPartition}.
314 tim 2700 \end{equation}
315 tim 2850 A canonical ensemble (NVT)is an ensemble of systems, each of which
316 tim 2700 can share its energy with a large heat reservoir. The distribution
317     of the total energy amongst the possible dynamical states is given
318     by the partition function,
319     \begin{equation}
320     \Omega (N,V,T) = e^{ - \beta A}
321     \label{introEquation:NVTPartition}
322     \end{equation}
323     Here, $A$ is the Helmholtz free energy which is defined as $ A = U -
324 tim 2819 TS$. Since most experiments are carried out under constant pressure
325 tim 2850 condition, the isothermal-isobaric ensemble (NPT) plays a very
326 tim 2819 important role in molecular simulations. The isothermal-isobaric
327     ensemble allow the system to exchange energy with a heat bath of
328     temperature $T$ and to change the volume as well. Its partition
329     function is given as
330 tim 2700 \begin{equation}
331     \Delta (N,P,T) = - e^{\beta G}.
332     \label{introEquation:NPTPartition}
333     \end{equation}
334     Here, $G = U - TS + PV$, and $G$ is called Gibbs free energy.
335    
336     \subsection{\label{introSection:liouville}Liouville's theorem}
337    
338 tim 2819 Liouville's theorem is the foundation on which statistical mechanics
339     rests. It describes the time evolution of the phase space
340 tim 2700 distribution function. In order to calculate the rate of change of
341 tim 2850 $\rho$, we begin from Eq.~\ref{introEquation:deltaN}. If we consider
342     the two faces perpendicular to the $q_1$ axis, which are located at
343     $q_1$ and $q_1 + \delta q_1$, the number of phase points leaving the
344     opposite face is given by the expression,
345 tim 2700 \begin{equation}
346     \left( {\rho + \frac{{\partial \rho }}{{\partial q_1 }}\delta q_1 }
347     \right)\left( {\dot q_1 + \frac{{\partial \dot q_1 }}{{\partial q_1
348     }}\delta q_1 } \right)\delta q_2 \ldots \delta q_f \delta p_1
349     \ldots \delta p_f .
350     \end{equation}
351     Summing all over the phase space, we obtain
352     \begin{equation}
353     \frac{{d(\delta N)}}{{dt}} = - \sum\limits_{i = 1}^f {\left[ {\rho
354     \left( {\frac{{\partial \dot q_i }}{{\partial q_i }} +
355     \frac{{\partial \dot p_i }}{{\partial p_i }}} \right) + \left(
356     {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i + \frac{{\partial
357     \rho }}{{\partial p_i }}\dot p_i } \right)} \right]} \delta q_1
358     \ldots \delta q_f \delta p_1 \ldots \delta p_f .
359     \end{equation}
360     Differentiating the equations of motion in Hamiltonian formalism
361     (\ref{introEquation:motionHamiltonianCoordinate},
362     \ref{introEquation:motionHamiltonianMomentum}), we can show,
363     \begin{equation}
364     \sum\limits_i {\left( {\frac{{\partial \dot q_i }}{{\partial q_i }}
365     + \frac{{\partial \dot p_i }}{{\partial p_i }}} \right)} = 0 ,
366     \end{equation}
367     which cancels the first terms of the right hand side. Furthermore,
368 tim 2819 dividing $ \delta q_1 \ldots \delta q_f \delta p_1 \ldots \delta
369 tim 2700 p_f $ in both sides, we can write out Liouville's theorem in a
370     simple form,
371     \begin{equation}
372     \frac{{\partial \rho }}{{\partial t}} + \sum\limits_{i = 1}^f
373     {\left( {\frac{{\partial \rho }}{{\partial q_i }}\dot q_i +
374     \frac{{\partial \rho }}{{\partial p_i }}\dot p_i } \right)} = 0 .
375     \label{introEquation:liouvilleTheorem}
376     \end{equation}
377    
378     Liouville's theorem states that the distribution function is
379     constant along any trajectory in phase space. In classical
380 tim 2850 statistical mechanics, since the number of members in an ensemble is
381     huge and constant, we can assume the local density has no reason
382     (other than classical mechanics) to change,
383 tim 2700 \begin{equation}
384     \frac{{\partial \rho }}{{\partial t}} = 0.
385     \label{introEquation:stationary}
386     \end{equation}
387     In such stationary system, the density of distribution $\rho$ can be
388     connected to the Hamiltonian $H$ through Maxwell-Boltzmann
389     distribution,
390     \begin{equation}
391     \rho \propto e^{ - \beta H}
392     \label{introEquation:densityAndHamiltonian}
393     \end{equation}
394    
395 tim 2819 \subsubsection{\label{introSection:phaseSpaceConservation}\textbf{Conservation of Phase Space}}
396 tim 2702 Lets consider a region in the phase space,
397     \begin{equation}
398     \delta v = \int { \ldots \int {dq_1 } ...dq_f dp_1 } ..dp_f .
399     \end{equation}
400     If this region is small enough, the density $\rho$ can be regarded
401 tim 2819 as uniform over the whole integral. Thus, the number of phase points
402     inside this region is given by,
403 tim 2702 \begin{equation}
404     \delta N = \rho \delta v = \rho \int { \ldots \int {dq_1 } ...dq_f
405     dp_1 } ..dp_f.
406     \end{equation}
407    
408     \begin{equation}
409     \frac{{d(\delta N)}}{{dt}} = \frac{{d\rho }}{{dt}}\delta v + \rho
410     \frac{d}{{dt}}(\delta v) = 0.
411     \end{equation}
412     With the help of stationary assumption
413     (\ref{introEquation:stationary}), we obtain the principle of the
414 tim 2819 \emph{conservation of volume in phase space},
415 tim 2702 \begin{equation}
416     \frac{d}{{dt}}(\delta v) = \frac{d}{{dt}}\int { \ldots \int {dq_1 }
417     ...dq_f dp_1 } ..dp_f = 0.
418     \label{introEquation:volumePreserving}
419     \end{equation}
420    
421 tim 2819 \subsubsection{\label{introSection:liouvilleInOtherForms}\textbf{Liouville's Theorem in Other Forms}}
422 tim 2702
423 tim 2700 Liouville's theorem can be expresses in a variety of different forms
424     which are convenient within different contexts. For any two function
425     $F$ and $G$ of the coordinates and momenta of a system, the Poisson
426     bracket ${F, G}$ is defined as
427     \begin{equation}
428     \left\{ {F,G} \right\} = \sum\limits_i {\left( {\frac{{\partial
429     F}}{{\partial q_i }}\frac{{\partial G}}{{\partial p_i }} -
430     \frac{{\partial F}}{{\partial p_i }}\frac{{\partial G}}{{\partial
431     q_i }}} \right)}.
432     \label{introEquation:poissonBracket}
433     \end{equation}
434     Substituting equations of motion in Hamiltonian formalism(
435 tim 2850 Eq.~\ref{introEquation:motionHamiltonianCoordinate} ,
436     Eq.~\ref{introEquation:motionHamiltonianMomentum} ) into
437     (Eq.~\ref{introEquation:liouvilleTheorem}), we can rewrite
438     Liouville's theorem using Poisson bracket notion,
439 tim 2700 \begin{equation}
440     \left( {\frac{{\partial \rho }}{{\partial t}}} \right) = - \left\{
441     {\rho ,H} \right\}.
442     \label{introEquation:liouvilleTheromInPoissin}
443     \end{equation}
444     Moreover, the Liouville operator is defined as
445     \begin{equation}
446     iL = \sum\limits_{i = 1}^f {\left( {\frac{{\partial H}}{{\partial
447     p_i }}\frac{\partial }{{\partial q_i }} - \frac{{\partial
448     H}}{{\partial q_i }}\frac{\partial }{{\partial p_i }}} \right)}
449     \label{introEquation:liouvilleOperator}
450     \end{equation}
451     In terms of Liouville operator, Liouville's equation can also be
452     expressed as
453     \begin{equation}
454     \left( {\frac{{\partial \rho }}{{\partial t}}} \right) = - iL\rho
455     \label{introEquation:liouvilleTheoremInOperator}
456     \end{equation}
457    
458 tim 2693 \subsection{\label{introSection:ergodic}The Ergodic Hypothesis}
459 tim 2692
460 tim 2695 Various thermodynamic properties can be calculated from Molecular
461     Dynamics simulation. By comparing experimental values with the
462     calculated properties, one can determine the accuracy of the
463 tim 2819 simulation and the quality of the underlying model. However, both
464     experiments and computer simulations are usually performed during a
465 tim 2695 certain time interval and the measurements are averaged over a
466     period of them which is different from the average behavior of
467 tim 2819 many-body system in Statistical Mechanics. Fortunately, the Ergodic
468     Hypothesis makes a connection between time average and the ensemble
469     average. It states that the time average and average over the
470 tim 2786 statistical ensemble are identical \cite{Frenkel1996, Leach2001}.
471 tim 2695 \begin{equation}
472 tim 2700 \langle A(q , p) \rangle_t = \mathop {\lim }\limits_{t \to \infty }
473     \frac{1}{t}\int\limits_0^t {A(q(t),p(t))dt = \int\limits_\Gamma
474     {A(q(t),p(t))} } \rho (q(t), p(t)) dqdp
475 tim 2695 \end{equation}
476 tim 2700 where $\langle A(q , p) \rangle_t$ is an equilibrium value of a
477     physical quantity and $\rho (p(t), q(t))$ is the equilibrium
478     distribution function. If an observation is averaged over a
479     sufficiently long time (longer than relaxation time), all accessible
480     microstates in phase space are assumed to be equally probed, giving
481     a properly weighted statistical average. This allows the researcher
482     freedom of choice when deciding how best to measure a given
483     observable. In case an ensemble averaged approach sounds most
484 tim 2786 reasonable, the Monte Carlo techniques\cite{Metropolis1949} can be
485 tim 2700 utilized. Or if the system lends itself to a time averaging
486     approach, the Molecular Dynamics techniques in
487     Sec.~\ref{introSection:molecularDynamics} will be the best
488     choice\cite{Frenkel1996}.
489 tim 2694
490 tim 2697 \section{\label{introSection:geometricIntegratos}Geometric Integrators}
491 tim 2819 A variety of numerical integrators have been proposed to simulate
492     the motions of atoms in MD simulation. They usually begin with
493     initial conditionals and move the objects in the direction governed
494     by the differential equations. However, most of them ignore the
495     hidden physical laws contained within the equations. Since 1990,
496     geometric integrators, which preserve various phase-flow invariants
497     such as symplectic structure, volume and time reversal symmetry, are
498     developed to address this issue\cite{Dullweber1997, McLachlan1998,
499 tim 2872 Leimkuhler1999}. The velocity Verlet method, which happens to be a
500 tim 2819 simple example of symplectic integrator, continues to gain
501     popularity in the molecular dynamics community. This fact can be
502     partly explained by its geometric nature.
503 tim 2697
504 tim 2819 \subsection{\label{introSection:symplecticManifold}Symplectic Manifolds}
505     A \emph{manifold} is an abstract mathematical space. It looks
506     locally like Euclidean space, but when viewed globally, it may have
507     more complicated structure. A good example of manifold is the
508     surface of Earth. It seems to be flat locally, but it is round if
509     viewed as a whole. A \emph{differentiable manifold} (also known as
510     \emph{smooth manifold}) is a manifold on which it is possible to
511     apply calculus on \emph{differentiable manifold}. A \emph{symplectic
512     manifold} is defined as a pair $(M, \omega)$ which consists of a
513 tim 2697 \emph{differentiable manifold} $M$ and a close, non-degenerated,
514     bilinear symplectic form, $\omega$. A symplectic form on a vector
515     space $V$ is a function $\omega(x, y)$ which satisfies
516     $\omega(\lambda_1x_1+\lambda_2x_2, y) = \lambda_1\omega(x_1, y)+
517     \lambda_2\omega(x_2, y)$, $\omega(x, y) = - \omega(y, x)$ and
518 tim 2819 $\omega(x, x) = 0$. The cross product operation in vector field is
519     an example of symplectic form.
520 tim 2697
521 tim 2819 One of the motivations to study \emph{symplectic manifolds} in
522 tim 2697 Hamiltonian Mechanics is that a symplectic manifold can represent
523     all possible configurations of the system and the phase space of the
524     system can be described by it's cotangent bundle. Every symplectic
525     manifold is even dimensional. For instance, in Hamilton equations,
526     coordinate and momentum always appear in pairs.
527    
528 tim 2698 \subsection{\label{introSection:ODE}Ordinary Differential Equations}
529 tim 2697
530 tim 2819 For an ordinary differential system defined as
531 tim 2698 \begin{equation}
532     \dot x = f(x)
533     \end{equation}
534 tim 2819 where $x = x(q,p)^T$, this system is a canonical Hamiltonian, if
535 tim 2698 \begin{equation}
536 tim 2699 f(r) = J\nabla _x H(r).
537 tim 2698 \end{equation}
538     $H = H (q, p)$ is Hamiltonian function and $J$ is the skew-symmetric
539     matrix
540     \begin{equation}
541     J = \left( {\begin{array}{*{20}c}
542     0 & I \\
543     { - I} & 0 \\
544     \end{array}} \right)
545     \label{introEquation:canonicalMatrix}
546     \end{equation}
547     where $I$ is an identity matrix. Using this notation, Hamiltonian
548     system can be rewritten as,
549     \begin{equation}
550     \frac{d}{{dt}}x = J\nabla _x H(x)
551     \label{introEquation:compactHamiltonian}
552     \end{equation}In this case, $f$ is
553     called a \emph{Hamiltonian vector field}.
554 tim 2697
555 tim 2789 Another generalization of Hamiltonian dynamics is Poisson
556     Dynamics\cite{Olver1986},
557 tim 2698 \begin{equation}
558     \dot x = J(x)\nabla _x H \label{introEquation:poissonHamiltonian}
559     \end{equation}
560     The most obvious change being that matrix $J$ now depends on $x$.
561    
562 tim 2702 \subsection{\label{introSection:exactFlow}Exact Flow}
563    
564 tim 2698 Let $x(t)$ be the exact solution of the ODE system,
565     \begin{equation}
566     \frac{{dx}}{{dt}} = f(x) \label{introEquation:ODE}
567     \end{equation}
568     The exact flow(solution) $\varphi_\tau$ is defined by
569     \[
570     x(t+\tau) =\varphi_\tau(x(t))
571     \]
572     where $\tau$ is a fixed time step and $\varphi$ is a map from phase
573 tim 2702 space to itself. The flow has the continuous group property,
574 tim 2698 \begin{equation}
575 tim 2702 \varphi _{\tau _1 } \circ \varphi _{\tau _2 } = \varphi _{\tau _1
576     + \tau _2 } .
577     \end{equation}
578     In particular,
579     \begin{equation}
580     \varphi _\tau \circ \varphi _{ - \tau } = I
581     \end{equation}
582     Therefore, the exact flow is self-adjoint,
583     \begin{equation}
584     \varphi _\tau = \varphi _{ - \tau }^{ - 1}.
585     \end{equation}
586     The exact flow can also be written in terms of the of an operator,
587     \begin{equation}
588     \varphi _\tau (x) = e^{\tau \sum\limits_i {f_i (x)\frac{\partial
589     }{{\partial x_i }}} } (x) \equiv \exp (\tau f)(x).
590     \label{introEquation:exponentialOperator}
591     \end{equation}
592    
593     In most cases, it is not easy to find the exact flow $\varphi_\tau$.
594 tim 2872 Instead, we use an approximate map, $\psi_\tau$, which is usually
595 tim 2702 called integrator. The order of an integrator $\psi_\tau$ is $p$, if
596     the Taylor series of $\psi_\tau$ agree to order $p$,
597     \begin{equation}
598 tim 2872 \psi_\tau(x) = x + \tau f(x) + O(\tau^{p+1})
599 tim 2698 \end{equation}
600    
601 tim 2702 \subsection{\label{introSection:geometricProperties}Geometric Properties}
602    
603 tim 2872 The hidden geometric properties\cite{Budd1999, Marsden1998} of an
604     ODE and its flow play important roles in numerical studies. Many of
605     them can be found in systems which occur naturally in applications.
606 tim 2702
607     Let $\varphi$ be the flow of Hamiltonian vector field, $\varphi$ is
608     a \emph{symplectic} flow if it satisfies,
609 tim 2698 \begin{equation}
610 tim 2703 {\varphi '}^T J \varphi ' = J.
611 tim 2698 \end{equation}
612     According to Liouville's theorem, the symplectic volume is invariant
613     under a Hamiltonian flow, which is the basis for classical
614 tim 2699 statistical mechanics. Furthermore, the flow of a Hamiltonian vector
615     field on a symplectic manifold can be shown to be a
616     symplectomorphism. As to the Poisson system,
617 tim 2698 \begin{equation}
618 tim 2703 {\varphi '}^T J \varphi ' = J \circ \varphi
619 tim 2698 \end{equation}
620 tim 2872 is the property that must be preserved by the integrator.
621 tim 2702
622     It is possible to construct a \emph{volume-preserving} flow for a
623 tim 2872 source free ODE ($ \nabla \cdot f = 0 $), if the flow satisfies $
624 tim 2702 \det d\varphi = 1$. One can show easily that a symplectic flow will
625     be volume-preserving.
626    
627 tim 2872 Changing the variables $y = h(x)$ in an ODE
628     (Eq.~\ref{introEquation:ODE}) will result in a new system,
629 tim 2698 \[
630     \dot y = \tilde f(y) = ((dh \cdot f)h^{ - 1} )(y).
631     \]
632     The vector filed $f$ has reversing symmetry $h$ if $f = - \tilde f$.
633     In other words, the flow of this vector field is reversible if and
634 tim 2702 only if $ h \circ \varphi ^{ - 1} = \varphi \circ h $.
635 tim 2698
636 tim 2705 A \emph{first integral}, or conserved quantity of a general
637     differential function is a function $ G:R^{2d} \to R^d $ which is
638     constant for all solutions of the ODE $\frac{{dx}}{{dt}} = f(x)$ ,
639     \[
640     \frac{{dG(x(t))}}{{dt}} = 0.
641     \]
642     Using chain rule, one may obtain,
643     \[
644     \sum\limits_i {\frac{{dG}}{{dx_i }}} f_i (x) = f \bullet \nabla G,
645     \]
646     which is the condition for conserving \emph{first integral}. For a
647     canonical Hamiltonian system, the time evolution of an arbitrary
648     smooth function $G$ is given by,
649 tim 2789
650     \begin{eqnarray}
651     \frac{{dG(x(t))}}{{dt}} & = & [\nabla _x G(x(t))]^T \dot x(t) \\
652     & = & [\nabla _x G(x(t))]^T J\nabla _x H(x(t)). \\
653 tim 2705 \label{introEquation:firstIntegral1}
654 tim 2789 \end{eqnarray}
655    
656    
657 tim 2705 Using poisson bracket notion, Equation
658     \ref{introEquation:firstIntegral1} can be rewritten as
659     \[
660     \frac{d}{{dt}}G(x(t)) = \left\{ {G,H} \right\}(x(t)).
661     \]
662     Therefore, the sufficient condition for $G$ to be the \emph{first
663     integral} of a Hamiltonian system is
664     \[
665     \left\{ {G,H} \right\} = 0.
666     \]
667     As well known, the Hamiltonian (or energy) H of a Hamiltonian system
668     is a \emph{first integral}, which is due to the fact $\{ H,H\} =
669     0$.
670    
671 tim 2789 When designing any numerical methods, one should always try to
672 tim 2702 preserve the structural properties of the original ODE and its flow.
673    
674 tim 2699 \subsection{\label{introSection:constructionSymplectic}Construction of Symplectic Methods}
675     A lot of well established and very effective numerical methods have
676     been successful precisely because of their symplecticities even
677     though this fact was not recognized when they were first
678 tim 2872 constructed. The most famous example is the Verlet-leapfrog method
679 tim 2819 in molecular dynamics. In general, symplectic integrators can be
680 tim 2699 constructed using one of four different methods.
681     \begin{enumerate}
682     \item Generating functions
683     \item Variational methods
684     \item Runge-Kutta methods
685     \item Splitting methods
686     \end{enumerate}
687 tim 2698
688 tim 2789 Generating function\cite{Channell1990} tends to lead to methods
689     which are cumbersome and difficult to use. In dissipative systems,
690     variational methods can capture the decay of energy
691     accurately\cite{Kane2000}. Since their geometrically unstable nature
692     against non-Hamiltonian perturbations, ordinary implicit Runge-Kutta
693     methods are not suitable for Hamiltonian system. Recently, various
694     high-order explicit Runge-Kutta methods
695     \cite{Owren1992,Chen2003}have been developed to overcome this
696 tim 2703 instability. However, due to computational penalty involved in
697 tim 2819 implementing the Runge-Kutta methods, they have not attracted much
698     attention from the Molecular Dynamics community. Instead, splitting
699     methods have been widely accepted since they exploit natural
700     decompositions of the system\cite{Tuckerman1992, McLachlan1998}.
701 tim 2702
702 tim 2819 \subsubsection{\label{introSection:splittingMethod}\textbf{Splitting Methods}}
703 tim 2702
704     The main idea behind splitting methods is to decompose the discrete
705     $\varphi_h$ as a composition of simpler flows,
706 tim 2699 \begin{equation}
707     \varphi _h = \varphi _{h_1 } \circ \varphi _{h_2 } \ldots \circ
708     \varphi _{h_n }
709     \label{introEquation:FlowDecomposition}
710     \end{equation}
711     where each of the sub-flow is chosen such that each represent a
712 tim 2702 simpler integration of the system.
713    
714     Suppose that a Hamiltonian system takes the form,
715     \[
716     H = H_1 + H_2.
717     \]
718     Here, $H_1$ and $H_2$ may represent different physical processes of
719     the system. For instance, they may relate to kinetic and potential
720     energy respectively, which is a natural decomposition of the
721     problem. If $H_1$ and $H_2$ can be integrated using exact flows
722     $\varphi_1(t)$ and $\varphi_2(t)$, respectively, a simple first
723 tim 2819 order expression is then given by the Lie-Trotter formula
724 tim 2699 \begin{equation}
725 tim 2702 \varphi _h = \varphi _{1,h} \circ \varphi _{2,h},
726     \label{introEquation:firstOrderSplitting}
727     \end{equation}
728     where $\varphi _h$ is the result of applying the corresponding
729     continuous $\varphi _i$ over a time $h$. By definition, as
730     $\varphi_i(t)$ is the exact solution of a Hamiltonian system, it
731     must follow that each operator $\varphi_i(t)$ is a symplectic map.
732     It is easy to show that any composition of symplectic flows yields a
733     symplectic map,
734     \begin{equation}
735 tim 2699 (\varphi '\phi ')^T J\varphi '\phi ' = \phi '^T \varphi '^T J\varphi
736 tim 2702 '\phi ' = \phi '^T J\phi ' = J,
737 tim 2699 \label{introEquation:SymplecticFlowComposition}
738     \end{equation}
739 tim 2702 where $\phi$ and $\psi$ both are symplectic maps. Thus operator
740     splitting in this context automatically generates a symplectic map.
741 tim 2699
742 tim 2702 The Lie-Trotter splitting(\ref{introEquation:firstOrderSplitting})
743     introduces local errors proportional to $h^2$, while Strang
744     splitting gives a second-order decomposition,
745     \begin{equation}
746     \varphi _h = \varphi _{1,h/2} \circ \varphi _{2,h} \circ \varphi
747 tim 2706 _{1,h/2} , \label{introEquation:secondOrderSplitting}
748 tim 2702 \end{equation}
749 tim 2819 which has a local error proportional to $h^3$. The Sprang
750     splitting's popularity in molecular simulation community attribute
751     to its symmetric property,
752 tim 2702 \begin{equation}
753     \varphi _h^{ - 1} = \varphi _{ - h}.
754 tim 2703 \label{introEquation:timeReversible}
755 tim 2844 \end{equation},appendixFig:architecture
756 tim 2702
757 tim 2872 \subsubsection{\label{introSection:exampleSplittingMethod}\textbf{Examples of the Splitting Method}}
758 tim 2702 The classical equation for a system consisting of interacting
759     particles can be written in Hamiltonian form,
760     \[
761     H = T + V
762     \]
763     where $T$ is the kinetic energy and $V$ is the potential energy.
764 tim 2872 Setting $H_1 = T, H_2 = V$ and applying the Strang splitting, one
765 tim 2702 obtains the following:
766     \begin{align}
767     q(\Delta t) &= q(0) + \dot{q}(0)\Delta t +
768     \frac{F[q(0)]}{m}\frac{\Delta t^2}{2}, %
769     \label{introEquation:Lp10a} \\%
770     %
771     \dot{q}(\Delta t) &= \dot{q}(0) + \frac{\Delta t}{2m}
772     \biggl [F[q(0)] + F[q(\Delta t)] \biggr]. %
773     \label{introEquation:Lp10b}
774     \end{align}
775     where $F(t)$ is the force at time $t$. This integration scheme is
776     known as \emph{velocity verlet} which is
777     symplectic(\ref{introEquation:SymplecticFlowComposition}),
778     time-reversible(\ref{introEquation:timeReversible}) and
779     volume-preserving (\ref{introEquation:volumePreserving}). These
780     geometric properties attribute to its long-time stability and its
781     popularity in the community. However, the most commonly used
782     velocity verlet integration scheme is written as below,
783     \begin{align}
784     \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) &=
785     \dot{q}(0) + \frac{\Delta t}{2m}\, F[q(0)], \label{introEquation:Lp9a}\\%
786     %
787     q(\Delta t) &= q(0) + \Delta t\, \dot{q}\biggl (\frac{\Delta t}{2}\biggr ),%
788     \label{introEquation:Lp9b}\\%
789     %
790     \dot{q}(\Delta t) &= \dot{q}\biggl (\frac{\Delta t}{2}\biggr ) +
791 tim 2872 \frac{\Delta t}{2m}\, F[q(t)]. \label{introEquation:Lp9c}
792 tim 2702 \end{align}
793     From the preceding splitting, one can see that the integration of
794     the equations of motion would follow:
795     \begin{enumerate}
796     \item calculate the velocities at the half step, $\frac{\Delta t}{2}$, from the forces calculated at the initial position.
797    
798     \item Use the half step velocities to move positions one whole step, $\Delta t$.
799    
800 tim 2872 \item Evaluate the forces at the new positions, $\mathbf{q}(\Delta t)$, and use the new forces to complete the velocity move.
801 tim 2702
802     \item Repeat from step 1 with the new position, velocities, and forces assuming the roles of the initial values.
803     \end{enumerate}
804    
805 tim 2872 By simply switching the order of the propagators in the splitting
806     and composing a new integrator, the \emph{position verlet}
807     integrator, can be generated,
808 tim 2702 \begin{align}
809     \dot q(\Delta t) &= \dot q(0) + \Delta tF(q(0))\left[ {q(0) +
810     \frac{{\Delta t}}{{2m}}\dot q(0)} \right], %
811     \label{introEquation:positionVerlet1} \\%
812     %
813 tim 2703 q(\Delta t) &= q(0) + \frac{{\Delta t}}{2}\left[ {\dot q(0) + \dot
814 tim 2702 q(\Delta t)} \right]. %
815 tim 2719 \label{introEquation:positionVerlet2}
816 tim 2702 \end{align}
817    
818 tim 2819 \subsubsection{\label{introSection:errorAnalysis}\textbf{Error Analysis and Higher Order Methods}}
819 tim 2702
820 tim 2872 The Baker-Campbell-Hausdorff formula can be used to determine the
821     local error of splitting method in terms of the commutator of the
822 tim 2702 operators(\ref{introEquation:exponentialOperator}) associated with
823 tim 2872 the sub-flow. For operators $hX$ and $hY$ which are associated with
824 tim 2726 $\varphi_1(t)$ and $\varphi_2(t)$ respectively , we have
825 tim 2702 \begin{equation}
826     \exp (hX + hY) = \exp (hZ)
827     \end{equation}
828     where
829     \begin{equation}
830     hZ = hX + hY + \frac{{h^2 }}{2}[X,Y] + \frac{{h^3 }}{2}\left(
831     {[X,[X,Y]] + [Y,[Y,X]]} \right) + \ldots .
832     \end{equation}
833     Here, $[X,Y]$ is the commutators of operator $X$ and $Y$ given by
834     \[
835     [X,Y] = XY - YX .
836     \]
837 tim 2872 Applying the Baker-Campbell-Hausdorff formula\cite{Varadarajan1974}
838     to the Sprang splitting, we can obtain
839 tim 2779 \begin{eqnarray*}
840 tim 2778 \exp (h X/2)\exp (h Y)\exp (h X/2) & = & \exp (h X + h Y + h^2 [X,Y]/4 + h^2 [Y,X]/4 \\
841     & & \mbox{} + h^2 [X,X]/8 + h^2 [Y,Y]/8 \\
842 tim 2779 & & \mbox{} + h^3 [Y,[Y,X]]/12 - h^3[X,[X,Y]]/24 + \ldots )
843     \end{eqnarray*}
844 tim 2872 Since \[ [X,Y] + [Y,X] = 0\] and \[ [X,X] = 0,\] the dominant local
845 tim 2702 error of Spring splitting is proportional to $h^3$. The same
846 tim 2872 procedure can be applied to a general splitting, of the form
847 tim 2702 \begin{equation}
848     \varphi _{b_m h}^2 \circ \varphi _{a_m h}^1 \circ \varphi _{b_{m -
849     1} h}^2 \circ \ldots \circ \varphi _{a_1 h}^1 .
850     \end{equation}
851 tim 2872 A careful choice of coefficient $a_1 \ldots b_m$ will lead to higher
852     order methods. Yoshida proposed an elegant way to compose higher
853 tim 2789 order methods based on symmetric splitting\cite{Yoshida1990}. Given
854     a symmetric second order base method $ \varphi _h^{(2)} $, a
855     fourth-order symmetric method can be constructed by composing,
856 tim 2702 \[
857     \varphi _h^{(4)} = \varphi _{\alpha h}^{(2)} \circ \varphi _{\beta
858     h}^{(2)} \circ \varphi _{\alpha h}^{(2)}
859     \]
860     where $ \alpha = - \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$ and $ \beta
861     = \frac{{2^{1/3} }}{{2 - 2^{1/3} }}$. Moreover, a symmetric
862     integrator $ \varphi _h^{(2n + 2)}$ can be composed by
863     \begin{equation}
864     \varphi _h^{(2n + 2)} = \varphi _{\alpha h}^{(2n)} \circ \varphi
865 tim 2872 _{\beta h}^{(2n)} \circ \varphi _{\alpha h}^{(2n)},
866 tim 2702 \end{equation}
867 tim 2872 if the weights are chosen as
868 tim 2702 \[
869     \alpha = - \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }},\beta =
870     \frac{{2^{1/(2n + 1)} }}{{2 - 2^{1/(2n + 1)} }} .
871     \]
872    
873 tim 2694 \section{\label{introSection:molecularDynamics}Molecular Dynamics}
874    
875 tim 2720 As one of the principal tools of molecular modeling, Molecular
876     dynamics has proven to be a powerful tool for studying the functions
877     of biological systems, providing structural, thermodynamic and
878     dynamical information. The basic idea of molecular dynamics is that
879     macroscopic properties are related to microscopic behavior and
880     microscopic behavior can be calculated from the trajectories in
881     simulations. For instance, instantaneous temperature of an
882     Hamiltonian system of $N$ particle can be measured by
883     \[
884 tim 2725 T = \sum\limits_{i = 1}^N {\frac{{m_i v_i^2 }}{{fk_B }}}
885 tim 2720 \]
886     where $m_i$ and $v_i$ are the mass and velocity of $i$th particle
887     respectively, $f$ is the number of degrees of freedom, and $k_B$ is
888     the boltzman constant.
889 tim 2694
890 tim 2720 A typical molecular dynamics run consists of three essential steps:
891     \begin{enumerate}
892     \item Initialization
893     \begin{enumerate}
894     \item Preliminary preparation
895     \item Minimization
896     \item Heating
897     \item Equilibration
898     \end{enumerate}
899     \item Production
900     \item Analysis
901     \end{enumerate}
902     These three individual steps will be covered in the following
903     sections. Sec.~\ref{introSec:initialSystemSettings} deals with the
904 tim 2801 initialization of a simulation. Sec.~\ref{introSection:production}
905 tim 2872 will discusse issues in production run.
906 tim 2801 Sec.~\ref{introSection:Analysis} provides the theoretical tools for
907     trajectory analysis.
908 tim 2719
909 tim 2720 \subsection{\label{introSec:initialSystemSettings}Initialization}
910 tim 2719
911 tim 2819 \subsubsection{\textbf{Preliminary preparation}}
912 tim 2719
913 tim 2720 When selecting the starting structure of a molecule for molecular
914     simulation, one may retrieve its Cartesian coordinates from public
915     databases, such as RCSB Protein Data Bank \textit{etc}. Although
916     thousands of crystal structures of molecules are discovered every
917     year, many more remain unknown due to the difficulties of
918 tim 2872 purification and crystallization. Even for molecules with known
919     structure, some important information is missing. For example, a
920 tim 2720 missing hydrogen atom which acts as donor in hydrogen bonding must
921     be added. Moreover, in order to include electrostatic interaction,
922     one may need to specify the partial charges for individual atoms.
923     Under some circumstances, we may even need to prepare the system in
924 tim 2872 a special configuration. For instance, when studying transport
925     phenomenon in membrane systems, we may prepare the lipids in a
926     bilayer structure instead of placing lipids randomly in solvent,
927     since we are not interested in the slow self-aggregation process.
928 tim 2694
929 tim 2819 \subsubsection{\textbf{Minimization}}
930 tim 2705
931 tim 2720 It is quite possible that some of molecules in the system from
932 tim 2872 preliminary preparation may be overlapping with each other. This
933     close proximity leads to high initial potential energy which
934     consequently jeopardizes any molecular dynamics simulations. To
935     remove these steric overlaps, one typically performs energy
936     minimization to find a more reasonable conformation. Several energy
937     minimization methods have been developed to exploit the energy
938     surface and to locate the local minimum. While converging slowly
939     near the minimum, steepest descent method is extremely robust when
940     systems are strongly anharmonic. Thus, it is often used to refine
941     structure from crystallographic data. Relied on the gradient or
942     hessian, advanced methods like Newton-Raphson converge rapidly to a
943     local minimum, but become unstable if the energy surface is far from
944     quadratic. Another factor that must be taken into account, when
945 tim 2720 choosing energy minimization method, is the size of the system.
946     Steepest descent and conjugate gradient can deal with models of any
947 tim 2872 size. Because of the limits on computer memory to store the hessian
948     matrix and the computing power needed to diagonalized these
949     matrices, most Newton-Raphson methods can not be used with very
950     large systems.
951 tim 2694
952 tim 2819 \subsubsection{\textbf{Heating}}
953 tim 2720
954     Typically, Heating is performed by assigning random velocities
955 tim 2872 according to a Maxwell-Boltzman distribution for a desired
956     temperature. Beginning at a lower temperature and gradually
957     increasing the temperature by assigning larger random velocities, we
958     end up with setting the temperature of the system to a final
959     temperature at which the simulation will be conducted. In heating
960     phase, we should also keep the system from drifting or rotating as a
961     whole. To do this, the net linear momentum and angular momentum of
962     the system is shifted to zero after each resampling from the Maxwell
963     -Boltzman distribution.
964 tim 2720
965 tim 2819 \subsubsection{\textbf{Equilibration}}
966 tim 2720
967     The purpose of equilibration is to allow the system to evolve
968     spontaneously for a period of time and reach equilibrium. The
969     procedure is continued until various statistical properties, such as
970     temperature, pressure, energy, volume and other structural
971     properties \textit{etc}, become independent of time. Strictly
972     speaking, minimization and heating are not necessary, provided the
973     equilibration process is long enough. However, these steps can serve
974     as a means to arrive at an equilibrated structure in an effective
975     way.
976    
977     \subsection{\label{introSection:production}Production}
978    
979 tim 2872 The production run is the most important step of the simulation, in
980 tim 2725 which the equilibrated structure is used as a starting point and the
981     motions of the molecules are collected for later analysis. In order
982     to capture the macroscopic properties of the system, the molecular
983 tim 2872 dynamics simulation must be performed by sampling correctly and
984     efficiently from the relevant thermodynamic ensemble.
985 tim 2720
986 tim 2725 The most expensive part of a molecular dynamics simulation is the
987     calculation of non-bonded forces, such as van der Waals force and
988     Coulombic forces \textit{etc}. For a system of $N$ particles, the
989     complexity of the algorithm for pair-wise interactions is $O(N^2 )$,
990     which making large simulations prohibitive in the absence of any
991 tim 2872 algorithmic tricks.
992 tim 2720
993 tim 2872 A natural approach to avoid system size issues is to represent the
994 tim 2725 bulk behavior by a finite number of the particles. However, this
995 tim 2872 approach will suffer from the surface effect at the edges of the
996     simulation. To offset this, \textit{Periodic boundary conditions}
997     (see Fig.~\ref{introFig:pbc}) is developed to simulate bulk
998     properties with a relatively small number of particles. In this
999     method, the simulation box is replicated throughout space to form an
1000     infinite lattice. During the simulation, when a particle moves in
1001     the primary cell, its image in other cells move in exactly the same
1002     direction with exactly the same orientation. Thus, as a particle
1003     leaves the primary cell, one of its images will enter through the
1004     opposite face.
1005 tim 2789 \begin{figure}
1006     \centering
1007     \includegraphics[width=\linewidth]{pbc.eps}
1008     \caption[An illustration of periodic boundary conditions]{A 2-D
1009     illustration of periodic boundary conditions. As one particle leaves
1010     the left of the simulation box, an image of it enters the right.}
1011     \label{introFig:pbc}
1012     \end{figure}
1013 tim 2725
1014     %cutoff and minimum image convention
1015     Another important technique to improve the efficiency of force
1016 tim 2872 evaluation is to apply spherical cutoff where particles farther than
1017     a predetermined distance are not included in the calculation
1018 tim 2725 \cite{Frenkel1996}. The use of a cutoff radius will cause a
1019 tim 2730 discontinuity in the potential energy curve. Fortunately, one can
1020 tim 2872 shift simple radial potential to ensure the potential curve go
1021     smoothly to zero at the cutoff radius. The cutoff strategy works
1022     well for Lennard-Jones interaction because of its short range
1023     nature. However, simply truncating the electrostatic interaction
1024     with the use of cutoffs has been shown to lead to severe artifacts
1025     in simulations. The Ewald summation, in which the slowly decaying
1026     Coulomb potential is transformed into direct and reciprocal sums
1027     with rapid and absolute convergence, has proved to minimize the
1028     periodicity artifacts in liquid simulations. Taking the advantages
1029     of the fast Fourier transform (FFT) for calculating discrete Fourier
1030     transforms, the particle mesh-based
1031 tim 2789 methods\cite{Hockney1981,Shimada1993, Luty1994} are accelerated from
1032 tim 2872 $O(N^{3/2})$ to $O(N logN)$. An alternative approach is the
1033     \emph{fast multipole method}\cite{Greengard1987, Greengard1994},
1034     which treats Coulombic interactions exactly at short range, and
1035     approximate the potential at long range through multipolar
1036     expansion. In spite of their wide acceptance at the molecular
1037     simulation community, these two methods are difficult to implement
1038     correctly and efficiently. Instead, we use a damped and
1039     charge-neutralized Coulomb potential method developed by Wolf and
1040     his coworkers\cite{Wolf1999}. The shifted Coulomb potential for
1041     particle $i$ and particle $j$ at distance $r_{rj}$ is given by:
1042 tim 2725 \begin{equation}
1043     V(r_{ij})= \frac{q_i q_j \textrm{erfc}(\alpha
1044     r_{ij})}{r_{ij}}-\lim_{r_{ij}\rightarrow
1045     R_\textrm{c}}\left\{\frac{q_iq_j \textrm{erfc}(\alpha
1046     r_{ij})}{r_{ij}}\right\}. \label{introEquation:shiftedCoulomb}
1047     \end{equation}
1048     where $\alpha$ is the convergence parameter. Due to the lack of
1049     inherent periodicity and rapid convergence,this method is extremely
1050     efficient and easy to implement.
1051 tim 2789 \begin{figure}
1052     \centering
1053     \includegraphics[width=\linewidth]{shifted_coulomb.eps}
1054     \caption[An illustration of shifted Coulomb potential]{An
1055     illustration of shifted Coulomb potential.}
1056     \label{introFigure:shiftedCoulomb}
1057     \end{figure}
1058 tim 2725
1059     %multiple time step
1060    
1061 tim 2720 \subsection{\label{introSection:Analysis} Analysis}
1062    
1063 tim 2872 Recently, advanced visualization technique have become applied to
1064 tim 2721 monitor the motions of molecules. Although the dynamics of the
1065     system can be described qualitatively from animation, quantitative
1066 tim 2872 trajectory analysis are more useful. According to the principles of
1067     Statistical Mechanics, Sec.~\ref{introSection:statisticalMechanics},
1068     one can compute thermodynamic properties, analyze fluctuations of
1069     structural parameters, and investigate time-dependent processes of
1070     the molecule from the trajectories.
1071 tim 2721
1072 tim 2872 \subsubsection{\label{introSection:thermodynamicsProperties}\textbf{Thermodynamic Properties}}
1073 tim 2721
1074 tim 2872 Thermodynamic properties, which can be expressed in terms of some
1075 tim 2725 function of the coordinates and momenta of all particles in the
1076     system, can be directly computed from molecular dynamics. The usual
1077     way to measure the pressure is based on virial theorem of Clausius
1078     which states that the virial is equal to $-3Nk_BT$. For a system
1079     with forces between particles, the total virial, $W$, contains the
1080     contribution from external pressure and interaction between the
1081     particles:
1082     \[
1083     W = - 3PV + \left\langle {\sum\limits_{i < j} {r{}_{ij} \cdot
1084     f_{ij} } } \right\rangle
1085     \]
1086     where $f_{ij}$ is the force between particle $i$ and $j$ at a
1087     distance $r_{ij}$. Thus, the expression for the pressure is given
1088     by:
1089     \begin{equation}
1090     P = \frac{{Nk_B T}}{V} - \frac{1}{{3V}}\left\langle {\sum\limits_{i
1091     < j} {r{}_{ij} \cdot f_{ij} } } \right\rangle
1092     \end{equation}
1093    
1094 tim 2819 \subsubsection{\label{introSection:structuralProperties}\textbf{Structural Properties}}
1095 tim 2721
1096     Structural Properties of a simple fluid can be described by a set of
1097 tim 2872 distribution functions. Among these functions,the \emph{pair
1098 tim 2721 distribution function}, also known as \emph{radial distribution
1099 tim 2872 function}, is of most fundamental importance to liquid theory.
1100     Experimentally, pair distribution function can be gathered by
1101     Fourier transforming raw data from a series of neutron diffraction
1102     experiments and integrating over the surface factor
1103     \cite{Powles1973}. The experimental results can serve as a criterion
1104     to justify the correctness of a liquid model. Moreover, various
1105     equilibrium thermodynamic and structural properties can also be
1106     expressed in terms of radial distribution function \cite{Allen1987}.
1107 tim 2721
1108 tim 2872 The pair distribution functions $g(r)$ gives the probability that a
1109 tim 2721 particle $i$ will be located at a distance $r$ from a another
1110     particle $j$ in the system
1111     \[
1112     g(r) = \frac{V}{{N^2 }}\left\langle {\sum\limits_i {\sum\limits_{j
1113 tim 2872 \ne i} {\delta (r - r_{ij} )} } } \right\rangle = \fract{\rho
1114     (r)}{\rho}.
1115 tim 2721 \]
1116     Note that the delta function can be replaced by a histogram in
1117     computer simulation. Figure
1118     \ref{introFigure:pairDistributionFunction} shows a typical pair
1119     distribution function for the liquid argon system. The occurrence of
1120     several peaks in the plot of $g(r)$ suggests that it is more likely
1121     to find particles at certain radial values than at others. This is a
1122     result of the attractive interaction at such distances. Because of
1123     the strong repulsive forces at short distance, the probability of
1124 tim 2872 locating particles at distances less than about 3.7{\AA} from each
1125 tim 2721 other is essentially zero.
1126    
1127     %\begin{figure}
1128     %\centering
1129     %\includegraphics[width=\linewidth]{pdf.eps}
1130     %\caption[Pair distribution function for the liquid argon
1131     %]{Pair distribution function for the liquid argon}
1132     %\label{introFigure:pairDistributionFunction}
1133     %\end{figure}
1134    
1135 tim 2819 \subsubsection{\label{introSection:timeDependentProperties}\textbf{Time-dependent
1136     Properties}}
1137 tim 2721
1138     Time-dependent properties are usually calculated using \emph{time
1139 tim 2872 correlation functions}, which correlate random variables $A$ and $B$
1140     at two different times,
1141 tim 2721 \begin{equation}
1142     C_{AB} (t) = \left\langle {A(t)B(0)} \right\rangle.
1143     \label{introEquation:timeCorrelationFunction}
1144     \end{equation}
1145     If $A$ and $B$ refer to same variable, this kind of correlation
1146 tim 2872 function is called an \emph{autocorrelation function}. One example
1147     of an auto correlation function is the velocity auto-correlation
1148     function which is directly related to transport properties of
1149     molecular liquids:
1150 tim 2725 \[
1151     D = \frac{1}{3}\int\limits_0^\infty {\left\langle {v(t) \cdot v(0)}
1152     \right\rangle } dt
1153     \]
1154 tim 2872 where $D$ is diffusion constant. Unlike the velocity autocorrelation
1155     function, which is averaging over time origins and over all the
1156     atoms, the dipole autocorrelation functions are calculated for the
1157     entire system. The dipole autocorrelation function is given by:
1158 tim 2725 \[
1159     c_{dipole} = \left\langle {u_{tot} (t) \cdot u_{tot} (t)}
1160     \right\rangle
1161     \]
1162     Here $u_{tot}$ is the net dipole of the entire system and is given
1163     by
1164     \[
1165     u_{tot} (t) = \sum\limits_i {u_i (t)}
1166     \]
1167     In principle, many time correlation functions can be related with
1168     Fourier transforms of the infrared, Raman, and inelastic neutron
1169     scattering spectra of molecular liquids. In practice, one can
1170     extract the IR spectrum from the intensity of dipole fluctuation at
1171     each frequency using the following relationship:
1172     \[
1173     \hat c_{dipole} (v) = \int_{ - \infty }^\infty {c_{dipole} (t)e^{ -
1174     i2\pi vt} dt}
1175     \]
1176 tim 2721
1177 tim 2693 \section{\label{introSection:rigidBody}Dynamics of Rigid Bodies}
1178 tim 2692
1179 tim 2705 Rigid bodies are frequently involved in the modeling of different
1180     areas, from engineering, physics, to chemistry. For example,
1181     missiles and vehicle are usually modeled by rigid bodies. The
1182     movement of the objects in 3D gaming engine or other physics
1183 tim 2872 simulator is governed by rigid body dynamics. In molecular
1184     simulations, rigid bodies are used to simplify protein-protein
1185     docking studies\cite{Gray2003}.
1186 tim 2694
1187 tim 2705 It is very important to develop stable and efficient methods to
1188 tim 2872 integrate the equations of motion for orientational degrees of
1189     freedom. Euler angles are the natural choice to describe the
1190     rotational degrees of freedom. However, due to $\frac {1}{sin
1191     \theta}$ singularities, the numerical integration of corresponding
1192     equations of motion is very inefficient and inaccurate. Although an
1193     alternative integrator using multiple sets of Euler angles can
1194     overcome this difficulty\cite{Barojas1973}, the computational
1195     penalty and the loss of angular momentum conservation still remain.
1196     A singularity-free representation utilizing quaternions was
1197     developed by Evans in 1977\cite{Evans1977}. Unfortunately, this
1198     approach uses a nonseparable Hamiltonian resulting from the
1199     quaternion representation, which prevents the symplectic algorithm
1200     to be utilized. Another different approach is to apply holonomic
1201     constraints to the atoms belonging to the rigid body. Each atom
1202     moves independently under the normal forces deriving from potential
1203     energy and constraint forces which are used to guarantee the
1204     rigidness. However, due to their iterative nature, the SHAKE and
1205     Rattle algorithms also converge very slowly when the number of
1206     constraints increases\cite{Ryckaert1977, Andersen1983}.
1207 tim 2694
1208 tim 2872 A break-through in geometric literature suggests that, in order to
1209 tim 2705 develop a long-term integration scheme, one should preserve the
1210 tim 2872 symplectic structure of the flow. By introducing a conjugate
1211     momentum to the rotation matrix $Q$ and re-formulating Hamiltonian's
1212     equation, a symplectic integrator, RSHAKE\cite{Kol1997}, was
1213     proposed to evolve the Hamiltonian system in a constraint manifold
1214     by iteratively satisfying the orthogonality constraint $Q^T Q = 1$.
1215     An alternative method using the quaternion representation was
1216     developed by Omelyan\cite{Omelyan1998}. However, both of these
1217     methods are iterative and inefficient. In this section, we descibe a
1218 tim 2789 symplectic Lie-Poisson integrator for rigid body developed by
1219     Dullweber and his coworkers\cite{Dullweber1997} in depth.
1220 tim 2705
1221 tim 2872 \subsection{\label{introSection:constrainedHamiltonianRB}Constrained Hamiltonian for Rigid Bodies}
1222     The motion of a rigid body is Hamiltonian with the Hamiltonian
1223 tim 2713 function
1224 tim 2706 \begin{equation}
1225     H = \frac{1}{2}(p^T m^{ - 1} p) + \frac{1}{2}tr(PJ^{ - 1} P) +
1226     V(q,Q) + \frac{1}{2}tr[(QQ^T - 1)\Lambda ].
1227     \label{introEquation:RBHamiltonian}
1228     \end{equation}
1229     Here, $q$ and $Q$ are the position and rotation matrix for the
1230     rigid-body, $p$ and $P$ are conjugate momenta to $q$ and $Q$ , and
1231     $J$, a diagonal matrix, is defined by
1232     \[
1233     I_{ii}^{ - 1} = \frac{1}{2}\sum\limits_{i \ne j} {J_{jj}^{ - 1} }
1234     \]
1235     where $I_{ii}$ is the diagonal element of the inertia tensor. This
1236 tim 2872 constrained Hamiltonian equation is subjected to a holonomic
1237     constraint,
1238 tim 2706 \begin{equation}
1239 tim 2726 Q^T Q = 1, \label{introEquation:orthogonalConstraint}
1240 tim 2706 \end{equation}
1241 tim 2872 which is used to ensure rotation matrix's unitarity. Differentiating
1242     \ref{introEquation:orthogonalConstraint} and using Equation
1243     \ref{introEquation:RBMotionMomentum}, one may obtain,
1244 tim 2706 \begin{equation}
1245 tim 2707 Q^T PJ^{ - 1} + J^{ - 1} P^T Q = 0 . \\
1246 tim 2706 \label{introEquation:RBFirstOrderConstraint}
1247     \end{equation}
1248    
1249     Using Equation (\ref{introEquation:motionHamiltonianCoordinate},
1250     \ref{introEquation:motionHamiltonianMomentum}), one can write down
1251     the equations of motion,
1252    
1253 tim 2796 \begin{eqnarray}
1254     \frac{{dq}}{{dt}} & = & \frac{p}{m} \label{introEquation:RBMotionPosition}\\
1255     \frac{{dp}}{{dt}} & = & - \nabla _q V(q,Q) \label{introEquation:RBMotionMomentum}\\
1256     \frac{{dQ}}{{dt}} & = & PJ^{ - 1} \label{introEquation:RBMotionRotation}\\
1257     \frac{{dP}}{{dt}} & = & - \nabla _Q V(q,Q) - 2Q\Lambda . \label{introEquation:RBMotionP}
1258     \end{eqnarray}
1259    
1260 tim 2707 In general, there are two ways to satisfy the holonomic constraints.
1261 tim 2872 We can use a constraint force provided by a Lagrange multiplier on
1262     the normal manifold to keep the motion on constraint space. Or we
1263     can simply evolve the system on the constraint manifold. These two
1264     methods have been proved to be equivalent. The holonomic constraint
1265     and equations of motions define a constraint manifold for rigid
1266     bodies
1267 tim 2707 \[
1268     M = \left\{ {(Q,P):Q^T Q = 1,Q^T PJ^{ - 1} + J^{ - 1} P^T Q = 0}
1269     \right\}.
1270     \]
1271 tim 2706
1272 tim 2707 Unfortunately, this constraint manifold is not the cotangent bundle
1273     $T_{\star}SO(3)$. However, it turns out that under symplectic
1274     transformation, the cotangent space and the phase space are
1275 tim 2872 diffeomorphic. By introducing
1276 tim 2706 \[
1277 tim 2707 \tilde Q = Q,\tilde P = \frac{1}{2}\left( {P - QP^T Q} \right),
1278 tim 2706 \]
1279 tim 2707 the mechanical system subject to a holonomic constraint manifold $M$
1280     can be re-formulated as a Hamiltonian system on the cotangent space
1281     \[
1282     T^* SO(3) = \left\{ {(\tilde Q,\tilde P):\tilde Q^T \tilde Q =
1283     1,\tilde Q^T \tilde PJ^{ - 1} + J^{ - 1} P^T \tilde Q = 0} \right\}
1284     \]
1285 tim 2706
1286 tim 2707 For a body fixed vector $X_i$ with respect to the center of mass of
1287     the rigid body, its corresponding lab fixed vector $X_0^{lab}$ is
1288     given as
1289     \begin{equation}
1290     X_i^{lab} = Q X_i + q.
1291     \end{equation}
1292     Therefore, potential energy $V(q,Q)$ is defined by
1293     \[
1294     V(q,Q) = V(Q X_0 + q).
1295     \]
1296 tim 2713 Hence, the force and torque are given by
1297 tim 2707 \[
1298 tim 2713 \nabla _q V(q,Q) = F(q,Q) = \sum\limits_i {F_i (q,Q)},
1299 tim 2707 \]
1300 tim 2713 and
1301 tim 2707 \[
1302     \nabla _Q V(q,Q) = F(q,Q)X_i^t
1303     \]
1304 tim 2713 respectively.
1305 tim 2695
1306 tim 2707 As a common choice to describe the rotation dynamics of the rigid
1307 tim 2872 body, the angular momentum on the body fixed frame $\Pi = Q^t P$ is
1308     introduced to rewrite the equations of motion,
1309 tim 2707 \begin{equation}
1310     \begin{array}{l}
1311     \mathop \Pi \limits^ \bullet = J^{ - 1} \Pi ^T \Pi + Q^T \sum\limits_i {F_i (q,Q)X_i^T } - \Lambda \\
1312     \mathop Q\limits^{{\rm{ }} \bullet } = Q\Pi {\rm{ }}J^{ - 1} \\
1313     \end{array}
1314     \label{introEqaution:RBMotionPI}
1315     \end{equation}
1316     , as well as holonomic constraints,
1317     \[
1318     \begin{array}{l}
1319     \Pi J^{ - 1} + J^{ - 1} \Pi ^t = 0 \\
1320     Q^T Q = 1 \\
1321     \end{array}
1322     \]
1323 tim 2692
1324 tim 2707 For a vector $v(v_1 ,v_2 ,v_3 ) \in R^3$ and a matrix $\hat v \in
1325     so(3)^ \star$, the hat-map isomorphism,
1326     \begin{equation}
1327     v(v_1 ,v_2 ,v_3 ) \Leftrightarrow \hat v = \left(
1328     {\begin{array}{*{20}c}
1329     0 & { - v_3 } & {v_2 } \\
1330     {v_3 } & 0 & { - v_1 } \\
1331     { - v_2 } & {v_1 } & 0 \\
1332     \end{array}} \right),
1333     \label{introEquation:hatmapIsomorphism}
1334     \end{equation}
1335     will let us associate the matrix products with traditional vector
1336     operations
1337     \[
1338     \hat vu = v \times u
1339     \]
1340     Using \ref{introEqaution:RBMotionPI}, one can construct a skew
1341     matrix,
1342     \begin{equation}
1343 tim 2797 (\mathop \Pi \limits^ \bullet - \mathop \Pi \limits^ {\bullet ^T}
1344 tim 2707 ){\rm{ }} = {\rm{ }}(\Pi - \Pi ^T ){\rm{ }}(J^{ - 1} \Pi + \Pi J^{
1345     - 1} ) + \sum\limits_i {[Q^T F_i (r,Q)X_i^T - X_i F_i (r,Q)^T Q]} -
1346     (\Lambda - \Lambda ^T ) . \label{introEquation:skewMatrixPI}
1347     \end{equation}
1348     Since $\Lambda$ is symmetric, the last term of Equation
1349 tim 2713 \ref{introEquation:skewMatrixPI} is zero, which implies the Lagrange
1350     multiplier $\Lambda$ is absent from the equations of motion. This
1351 tim 2872 unique property eliminates the requirement of iterations which can
1352 tim 2789 not be avoided in other methods\cite{Kol1997, Omelyan1998}.
1353 tim 2707
1354 tim 2872 Applying the hat-map isomorphism, we obtain the equation of motion
1355     for angular momentum on body frame
1356 tim 2713 \begin{equation}
1357     \dot \pi = \pi \times I^{ - 1} \pi + \sum\limits_i {\left( {Q^T
1358     F_i (r,Q)} \right) \times X_i }.
1359     \label{introEquation:bodyAngularMotion}
1360     \end{equation}
1361 tim 2707 In the same manner, the equation of motion for rotation matrix is
1362     given by
1363     \[
1364 tim 2713 \dot Q = Qskew(I^{ - 1} \pi )
1365 tim 2707 \]
1366    
1367 tim 2713 \subsection{\label{introSection:SymplecticFreeRB}Symplectic
1368     Lie-Poisson Integrator for Free Rigid Body}
1369 tim 2707
1370 tim 2872 If there are no external forces exerted on the rigid body, the only
1371     contribution to the rotational motion is from the kinetic energy
1372     (the first term of \ref{introEquation:bodyAngularMotion}). The free
1373     rigid body is an example of a Lie-Poisson system with Hamiltonian
1374     function
1375 tim 2713 \begin{equation}
1376     T^r (\pi ) = T_1 ^r (\pi _1 ) + T_2^r (\pi _2 ) + T_3^r (\pi _3 )
1377     \label{introEquation:rotationalKineticRB}
1378     \end{equation}
1379     where $T_i^r (\pi _i ) = \frac{{\pi _i ^2 }}{{2I_i }}$ and
1380     Lie-Poisson structure matrix,
1381     \begin{equation}
1382     J(\pi ) = \left( {\begin{array}{*{20}c}
1383     0 & {\pi _3 } & { - \pi _2 } \\
1384     { - \pi _3 } & 0 & {\pi _1 } \\
1385     {\pi _2 } & { - \pi _1 } & 0 \\
1386     \end{array}} \right)
1387     \end{equation}
1388     Thus, the dynamics of free rigid body is governed by
1389     \begin{equation}
1390     \frac{d}{{dt}}\pi = J(\pi )\nabla _\pi T^r (\pi )
1391     \end{equation}
1392 tim 2707
1393 tim 2713 One may notice that each $T_i^r$ in Equation
1394     \ref{introEquation:rotationalKineticRB} can be solved exactly. For
1395     instance, the equations of motion due to $T_1^r$ are given by
1396     \begin{equation}
1397     \frac{d}{{dt}}\pi = R_1 \pi ,\frac{d}{{dt}}Q = QR_1
1398     \label{introEqaution:RBMotionSingleTerm}
1399     \end{equation}
1400     where
1401     \[ R_1 = \left( {\begin{array}{*{20}c}
1402     0 & 0 & 0 \\
1403     0 & 0 & {\pi _1 } \\
1404     0 & { - \pi _1 } & 0 \\
1405     \end{array}} \right).
1406     \]
1407     The solutions of Equation \ref{introEqaution:RBMotionSingleTerm} is
1408 tim 2707 \[
1409 tim 2713 \pi (\Delta t) = e^{\Delta tR_1 } \pi (0),Q(\Delta t) =
1410     Q(0)e^{\Delta tR_1 }
1411 tim 2707 \]
1412 tim 2713 with
1413 tim 2707 \[
1414 tim 2713 e^{\Delta tR_1 } = \left( {\begin{array}{*{20}c}
1415     0 & 0 & 0 \\
1416     0 & {\cos \theta _1 } & {\sin \theta _1 } \\
1417     0 & { - \sin \theta _1 } & {\cos \theta _1 } \\
1418     \end{array}} \right),\theta _1 = \frac{{\pi _1 }}{{I_1 }}\Delta t.
1419 tim 2707 \]
1420 tim 2719 To reduce the cost of computing expensive functions in $e^{\Delta
1421 tim 2872 tR_1 }$, we can use Cayley transformation to obtain a single-aixs
1422     propagator,
1423 tim 2713 \[
1424     e^{\Delta tR_1 } \approx (1 - \Delta tR_1 )^{ - 1} (1 + \Delta tR_1
1425     )
1426     \]
1427 tim 2720 The flow maps for $T_2^r$ and $T_3^r$ can be found in the same
1428 tim 2872 manner. In order to construct a second-order symplectic method, we
1429     split the angular kinetic Hamiltonian function can into five terms
1430 tim 2707 \[
1431 tim 2713 T^r (\pi ) = \frac{1}{2}T_1 ^r (\pi _1 ) + \frac{1}{2}T_2^r (\pi _2
1432     ) + T_3^r (\pi _3 ) + \frac{1}{2}T_2^r (\pi _2 ) + \frac{1}{2}T_1 ^r
1433 tim 2872 (\pi _1 ).
1434     \]
1435     By concatenating the propagators corresponding to these five terms,
1436     we can obtain an symplectic integrator,
1437 tim 2713 \[
1438     \varphi _{\Delta t,T^r } = \varphi _{\Delta t/2,\pi _1 } \circ
1439 tim 2707 \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t,\pi _3 }
1440     \circ \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t/2,\pi
1441 tim 2713 _1 }.
1442 tim 2707 \]
1443    
1444 tim 2713 The non-canonical Lie-Poisson bracket ${F, G}$ of two function
1445     $F(\pi )$ and $G(\pi )$ is defined by
1446 tim 2707 \[
1447 tim 2713 \{ F,G\} (\pi ) = [\nabla _\pi F(\pi )]^T J(\pi )\nabla _\pi G(\pi
1448     )
1449     \]
1450     If the Poisson bracket of a function $F$ with an arbitrary smooth
1451     function $G$ is zero, $F$ is a \emph{Casimir}, which is the
1452     conserved quantity in Poisson system. We can easily verify that the
1453     norm of the angular momentum, $\parallel \pi
1454     \parallel$, is a \emph{Casimir}. Let$ F(\pi ) = S(\frac{{\parallel
1455     \pi \parallel ^2 }}{2})$ for an arbitrary function $ S:R \to R$ ,
1456     then by the chain rule
1457     \[
1458     \nabla _\pi F(\pi ) = S'(\frac{{\parallel \pi \parallel ^2
1459     }}{2})\pi
1460     \]
1461     Thus $ [\nabla _\pi F(\pi )]^T J(\pi ) = - S'(\frac{{\parallel \pi
1462     \parallel ^2 }}{2})\pi \times \pi = 0 $. This explicit
1463 tim 2872 Lie-Poisson integrator is found to be both extremely efficient and
1464     stable. These properties can be explained by the fact the small
1465     angle approximation is used and the norm of the angular momentum is
1466     conserved.
1467 tim 2713
1468     \subsection{\label{introSection:RBHamiltonianSplitting} Hamiltonian
1469     Splitting for Rigid Body}
1470    
1471     The Hamiltonian of rigid body can be separated in terms of kinetic
1472     energy and potential energy,
1473     \[
1474     H = T(p,\pi ) + V(q,Q)
1475     \]
1476     The equations of motion corresponding to potential energy and
1477     kinetic energy are listed in the below table,
1478 tim 2776 \begin{table}
1479     \caption{Equations of motion due to Potential and Kinetic Energies}
1480 tim 2713 \begin{center}
1481     \begin{tabular}{|l|l|}
1482     \hline
1483     % after \\: \hline or \cline{col1-col2} \cline{col3-col4} ...
1484     Potential & Kinetic \\
1485     $\frac{{dq}}{{dt}} = \frac{p}{m}$ & $\frac{d}{{dt}}q = p$ \\
1486     $\frac{d}{{dt}}p = - \frac{{\partial V}}{{\partial q}}$ & $ \frac{d}{{dt}}p = 0$ \\
1487     $\frac{d}{{dt}}Q = 0$ & $ \frac{d}{{dt}}Q = Qskew(I^{ - 1} j)$ \\
1488     $ \frac{d}{{dt}}\pi = \sum\limits_i {\left( {Q^T F_i (r,Q)} \right) \times X_i }$ & $\frac{d}{{dt}}\pi = \pi \times I^{ - 1} \pi$\\
1489     \hline
1490     \end{tabular}
1491     \end{center}
1492 tim 2776 \end{table}
1493 tim 2872 A second-order symplectic method is now obtained by the composition
1494     of the position and velocity propagators,
1495 tim 2713 \[
1496     \varphi _{\Delta t} = \varphi _{\Delta t/2,V} \circ \varphi
1497     _{\Delta t,T} \circ \varphi _{\Delta t/2,V}.
1498     \]
1499 tim 2719 Moreover, $\varphi _{\Delta t/2,V}$ can be divided into two
1500 tim 2872 sub-propagators which corresponding to force and torque
1501     respectively,
1502 tim 2713 \[
1503 tim 2707 \varphi _{\Delta t/2,V} = \varphi _{\Delta t/2,F} \circ \varphi
1504 tim 2713 _{\Delta t/2,\tau }.
1505 tim 2707 \]
1506 tim 2713 Since the associated operators of $\varphi _{\Delta t/2,F} $ and
1507 tim 2872 $\circ \varphi _{\Delta t/2,\tau }$ commute, the composition order
1508     inside $\varphi _{\Delta t/2,V}$ does not matter. Furthermore, the
1509     kinetic energy can be separated to translational kinetic term, $T^t
1510     (p)$, and rotational kinetic term, $T^r (\pi )$,
1511 tim 2713 \begin{equation}
1512     T(p,\pi ) =T^t (p) + T^r (\pi ).
1513     \end{equation}
1514     where $ T^t (p) = \frac{1}{2}p^T m^{ - 1} p $ and $T^r (\pi )$ is
1515     defined by \ref{introEquation:rotationalKineticRB}. Therefore, the
1516 tim 2872 corresponding propagators are given by
1517 tim 2713 \[
1518     \varphi _{\Delta t,T} = \varphi _{\Delta t,T^t } \circ \varphi
1519     _{\Delta t,T^r }.
1520     \]
1521 tim 2872 Finally, we obtain the overall symplectic propagators for freely
1522     moving rigid bodies
1523 tim 2713 \begin{equation}
1524     \begin{array}{c}
1525     \varphi _{\Delta t} = \varphi _{\Delta t/2,F} \circ \varphi _{\Delta t/2,\tau } \\
1526     \circ \varphi _{\Delta t,T^t } \circ \varphi _{\Delta t/2,\pi _1 } \circ \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t,\pi _3 } \circ \varphi _{\Delta t/2,\pi _2 } \circ \varphi _{\Delta t/2,\pi _1 } \\
1527     \circ \varphi _{\Delta t/2,\tau } \circ \varphi _{\Delta t/2,F} .\\
1528     \end{array}
1529     \label{introEquation:overallRBFlowMaps}
1530     \end{equation}
1531 tim 2707
1532 tim 2685 \section{\label{introSection:langevinDynamics}Langevin Dynamics}
1533 tim 2716 As an alternative to newtonian dynamics, Langevin dynamics, which
1534     mimics a simple heat bath with stochastic and dissipative forces,
1535     has been applied in a variety of studies. This section will review
1536 tim 2872 the theory of Langevin dynamics. A brief derivation of generalized
1537     Langevin equation will be given first. Following that, we will
1538     discuss the physical meaning of the terms appearing in the equation
1539     as well as the calculation of friction tensor from hydrodynamics
1540     theory.
1541 tim 2685
1542 tim 2719 \subsection{\label{introSection:generalizedLangevinDynamics}Derivation of Generalized Langevin Equation}
1543 tim 2685
1544 tim 2872 A harmonic bath model, in which an effective set of harmonic
1545 tim 2719 oscillators are used to mimic the effect of a linearly responding
1546     environment, has been widely used in quantum chemistry and
1547     statistical mechanics. One of the successful applications of
1548 tim 2872 Harmonic bath model is the derivation of the Generalized Langevin
1549     Dynamics (GLE). Lets consider a system, in which the degree of
1550 tim 2719 freedom $x$ is assumed to couple to the bath linearly, giving a
1551     Hamiltonian of the form
1552 tim 2696 \begin{equation}
1553     H = \frac{{p^2 }}{{2m}} + U(x) + H_B + \Delta U(x,x_1 , \ldots x_N)
1554 tim 2719 \label{introEquation:bathGLE}.
1555 tim 2696 \end{equation}
1556 tim 2872 Here $p$ is a momentum conjugate to $x$, $m$ is the mass associated
1557     with this degree of freedom, $H_B$ is a harmonic bath Hamiltonian,
1558 tim 2696 \[
1559 tim 2719 H_B = \sum\limits_{\alpha = 1}^N {\left\{ {\frac{{p_\alpha ^2
1560     }}{{2m_\alpha }} + \frac{1}{2}m_\alpha \omega _\alpha ^2 }
1561     \right\}}
1562 tim 2696 \]
1563 tim 2719 where the index $\alpha$ runs over all the bath degrees of freedom,
1564     $\omega _\alpha$ are the harmonic bath frequencies, $m_\alpha$ are
1565 tim 2872 the harmonic bath masses, and $\Delta U$ is a bilinear system-bath
1566 tim 2719 coupling,
1567 tim 2696 \[
1568     \Delta U = - \sum\limits_{\alpha = 1}^N {g_\alpha x_\alpha x}
1569     \]
1570 tim 2872 where $g_\alpha$ are the coupling constants between the bath
1571     coordinates ($x_ \apha$) and the system coordinate ($x$).
1572     Introducing
1573 tim 2696 \[
1574 tim 2719 W(x) = U(x) - \sum\limits_{\alpha = 1}^N {\frac{{g_\alpha ^2
1575     }}{{2m_\alpha w_\alpha ^2 }}} x^2
1576     \] and combining the last two terms in Equation
1577     \ref{introEquation:bathGLE}, we may rewrite the Harmonic bath
1578     Hamiltonian as
1579 tim 2696 \[
1580     H = \frac{{p^2 }}{{2m}} + W(x) + \sum\limits_{\alpha = 1}^N
1581     {\left\{ {\frac{{p_\alpha ^2 }}{{2m_\alpha }} + \frac{1}{2}m_\alpha
1582     w_\alpha ^2 \left( {x_\alpha - \frac{{g_\alpha }}{{m_\alpha
1583     w_\alpha ^2 }}x} \right)^2 } \right\}}
1584     \]
1585     Since the first two terms of the new Hamiltonian depend only on the
1586     system coordinates, we can get the equations of motion for
1587 tim 2872 Generalized Langevin Dynamics by Hamilton's equations,
1588 tim 2719 \begin{equation}
1589     m\ddot x = - \frac{{\partial W(x)}}{{\partial x}} -
1590     \sum\limits_{\alpha = 1}^N {g_\alpha \left( {x_\alpha -
1591     \frac{{g_\alpha }}{{m_\alpha w_\alpha ^2 }}x} \right)},
1592     \label{introEquation:coorMotionGLE}
1593     \end{equation}
1594     and
1595     \begin{equation}
1596     m\ddot x_\alpha = - m_\alpha w_\alpha ^2 \left( {x_\alpha -
1597     \frac{{g_\alpha }}{{m_\alpha w_\alpha ^2 }}x} \right).
1598     \label{introEquation:bathMotionGLE}
1599     \end{equation}
1600 tim 2696
1601 tim 2719 In order to derive an equation for $x$, the dynamics of the bath
1602     variables $x_\alpha$ must be solved exactly first. As an integral
1603     transform which is particularly useful in solving linear ordinary
1604 tim 2872 differential equations,the Laplace transform is the appropriate tool
1605     to solve this problem. The basic idea is to transform the difficult
1606 tim 2719 differential equations into simple algebra problems which can be
1607 tim 2872 solved easily. Then, by applying the inverse Laplace transform, also
1608     known as the Bromwich integral, we can retrieve the solutions of the
1609 tim 2719 original problems.
1610 tim 2696
1611 tim 2719 Let $f(t)$ be a function defined on $ [0,\infty ) $. The Laplace
1612     transform of f(t) is a new function defined as
1613 tim 2696 \[
1614 tim 2719 L(f(t)) \equiv F(p) = \int_0^\infty {f(t)e^{ - pt} dt}
1615 tim 2696 \]
1616 tim 2719 where $p$ is real and $L$ is called the Laplace Transform
1617     Operator. Below are some important properties of Laplace transform
1618 tim 2696
1619 tim 2789 \begin{eqnarray*}
1620     L(x + y) & = & L(x) + L(y) \\
1621     L(ax) & = & aL(x) \\
1622     L(\dot x) & = & pL(x) - px(0) \\
1623     L(\ddot x)& = & p^2 L(x) - px(0) - \dot x(0) \\
1624     L\left( {\int_0^t {g(t - \tau )h(\tau )d\tau } } \right)& = & G(p)H(p) \\
1625     \end{eqnarray*}
1626    
1627    
1628 tim 2872 Applying the Laplace transform to the bath coordinates, we obtain
1629 tim 2789 \begin{eqnarray*}
1630     p^2 L(x_\alpha ) - px_\alpha (0) - \dot x_\alpha (0) & = & - \omega _\alpha ^2 L(x_\alpha ) + \frac{{g_\alpha }}{{\omega _\alpha }}L(x) \\
1631     L(x_\alpha ) & = & \frac{{\frac{{g_\alpha }}{{\omega _\alpha }}L(x) + px_\alpha (0) + \dot x_\alpha (0)}}{{p^2 + \omega _\alpha ^2 }} \\
1632     \end{eqnarray*}
1633    
1634 tim 2719 By the same way, the system coordinates become
1635 tim 2789 \begin{eqnarray*}
1636     mL(\ddot x) & = & - \frac{1}{p}\frac{{\partial W(x)}}{{\partial x}} \\
1637     & & \mbox{} - \sum\limits_{\alpha = 1}^N {\left\{ { - \frac{{g_\alpha ^2 }}{{m_\alpha \omega _\alpha ^2 }}\frac{p}{{p^2 + \omega _\alpha ^2 }}pL(x) - \frac{p}{{p^2 + \omega _\alpha ^2 }}g_\alpha x_\alpha (0) - \frac{1}{{p^2 + \omega _\alpha ^2 }}g_\alpha \dot x_\alpha (0)} \right\}} \\
1638     \end{eqnarray*}
1639 tim 2696
1640 tim 2719 With the help of some relatively important inverse Laplace
1641     transformations:
1642 tim 2696 \[
1643 tim 2719 \begin{array}{c}
1644     L(\cos at) = \frac{p}{{p^2 + a^2 }} \\
1645     L(\sin at) = \frac{a}{{p^2 + a^2 }} \\
1646     L(1) = \frac{1}{p} \\
1647     \end{array}
1648 tim 2696 \]
1649 tim 2719 , we obtain
1650 tim 2794 \begin{eqnarray*}
1651     m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} -
1652 tim 2696 \sum\limits_{\alpha = 1}^N {\left\{ {\left( { - \frac{{g_\alpha ^2
1653     }}{{m_\alpha \omega _\alpha ^2 }}} \right)\int_0^t {\cos (\omega
1654 tim 2794 _\alpha t)\dot x(t - \tau )d\tau } } \right\}} \\
1655     & & + \sum\limits_{\alpha = 1}^N {\left\{ {\left[ {g_\alpha
1656     x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha }}}
1657     \right]\cos (\omega _\alpha t) + \frac{{g_\alpha \dot x_\alpha
1658     (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t)} \right\}}
1659     \end{eqnarray*}
1660     \begin{eqnarray*}
1661     m\ddot x & = & - \frac{{\partial W(x)}}{{\partial x}} - \int_0^t
1662 tim 2696 {\sum\limits_{\alpha = 1}^N {\left( { - \frac{{g_\alpha ^2
1663     }}{{m_\alpha \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha
1664 tim 2794 t)\dot x(t - \tau )d} \tau } \\
1665     & & + \sum\limits_{\alpha = 1}^N {\left\{ {\left[ {g_\alpha
1666     x_\alpha (0) - \frac{{g_\alpha }}{{m_\alpha \omega _\alpha }}}
1667     \right]\cos (\omega _\alpha t) + \frac{{g_\alpha \dot x_\alpha
1668     (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t)} \right\}}
1669     \end{eqnarray*}
1670 tim 2719 Introducing a \emph{dynamic friction kernel}
1671 tim 2696 \begin{equation}
1672 tim 2719 \xi (t) = \sum\limits_{\alpha = 1}^N {\left( { - \frac{{g_\alpha ^2
1673     }}{{m_\alpha \omega _\alpha ^2 }}} \right)\cos (\omega _\alpha t)}
1674     \label{introEquation:dynamicFrictionKernelDefinition}
1675     \end{equation}
1676     and \emph{a random force}
1677     \begin{equation}
1678     R(t) = \sum\limits_{\alpha = 1}^N {\left( {g_\alpha x_\alpha (0)
1679     - \frac{{g_\alpha ^2 }}{{m_\alpha \omega _\alpha ^2 }}x(0)}
1680     \right)\cos (\omega _\alpha t)} + \frac{{\dot x_\alpha
1681     (0)}}{{\omega _\alpha }}\sin (\omega _\alpha t),
1682     \label{introEquation:randomForceDefinition}
1683     \end{equation}
1684     the equation of motion can be rewritten as
1685     \begin{equation}
1686 tim 2696 m\ddot x = - \frac{{\partial W}}{{\partial x}} - \int_0^t {\xi
1687     (t)\dot x(t - \tau )d\tau } + R(t)
1688     \label{introEuqation:GeneralizedLangevinDynamics}
1689     \end{equation}
1690 tim 2719 which is known as the \emph{generalized Langevin equation}.
1691    
1692 tim 2819 \subsubsection{\label{introSection:randomForceDynamicFrictionKernel}\textbf{Random Force and Dynamic Friction Kernel}}
1693 tim 2719
1694     One may notice that $R(t)$ depends only on initial conditions, which
1695     implies it is completely deterministic within the context of a
1696     harmonic bath. However, it is easy to verify that $R(t)$ is totally
1697     uncorrelated to $x$ and $\dot x$,
1698 tim 2696 \[
1699 tim 2719 \begin{array}{l}
1700     \left\langle {x(t)R(t)} \right\rangle = 0, \\
1701     \left\langle {\dot x(t)R(t)} \right\rangle = 0. \\
1702     \end{array}
1703 tim 2696 \]
1704 tim 2719 This property is what we expect from a truly random process. As long
1705 tim 2872 as the model chosen for $R(t)$ was a gaussian distribution in
1706     general, the stochastic nature of the GLE still remains.
1707 tim 2696
1708 tim 2719 %dynamic friction kernel
1709     The convolution integral
1710 tim 2696 \[
1711 tim 2719 \int_0^t {\xi (t)\dot x(t - \tau )d\tau }
1712 tim 2696 \]
1713 tim 2719 depends on the entire history of the evolution of $x$, which implies
1714     that the bath retains memory of previous motions. In other words,
1715     the bath requires a finite time to respond to change in the motion
1716     of the system. For a sluggish bath which responds slowly to changes
1717     in the system coordinate, we may regard $\xi(t)$ as a constant
1718     $\xi(t) = \Xi_0$. Hence, the convolution integral becomes
1719     \[
1720     \int_0^t {\xi (t)\dot x(t - \tau )d\tau } = \xi _0 (x(t) - x(0))
1721     \]
1722     and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1723     \[
1724     m\ddot x = - \frac{\partial }{{\partial x}}\left( {W(x) +
1725     \frac{1}{2}\xi _0 (x - x_0 )^2 } \right) + R(t),
1726     \]
1727 tim 2872 which can be used to describe the effect of dynamic caging in
1728     viscous solvents. The other extreme is the bath that responds
1729     infinitely quickly to motions in the system. Thus, $\xi (t)$ can be
1730     taken as a $delta$ function in time:
1731 tim 2719 \[
1732     \xi (t) = 2\xi _0 \delta (t)
1733     \]
1734     Hence, the convolution integral becomes
1735     \[
1736     \int_0^t {\xi (t)\dot x(t - \tau )d\tau } = 2\xi _0 \int_0^t
1737     {\delta (t)\dot x(t - \tau )d\tau } = \xi _0 \dot x(t),
1738     \]
1739     and Equation \ref{introEuqation:GeneralizedLangevinDynamics} becomes
1740     \begin{equation}
1741     m\ddot x = - \frac{{\partial W(x)}}{{\partial x}} - \xi _0 \dot
1742     x(t) + R(t) \label{introEquation:LangevinEquation}
1743     \end{equation}
1744     which is known as the Langevin equation. The static friction
1745     coefficient $\xi _0$ can either be calculated from spectral density
1746 tim 2850 or be determined by Stokes' law for regular shaped particles. A
1747 tim 2719 briefly review on calculating friction tensor for arbitrary shaped
1748 tim 2720 particles is given in Sec.~\ref{introSection:frictionTensor}.
1749 tim 2696
1750 tim 2819 \subsubsection{\label{introSection:secondFluctuationDissipation}\textbf{The Second Fluctuation Dissipation Theorem}}
1751 tim 2719
1752     Defining a new set of coordinates,
1753 tim 2696 \[
1754     q_\alpha (t) = x_\alpha (t) - \frac{1}{{m_\alpha \omega _\alpha
1755     ^2 }}x(0)
1756 tim 2719 \],
1757     we can rewrite $R(T)$ as
1758 tim 2696 \[
1759 tim 2719 R(t) = \sum\limits_{\alpha = 1}^N {g_\alpha q_\alpha (t)}.
1760 tim 2696 \]
1761     And since the $q$ coordinates are harmonic oscillators,
1762 tim 2789
1763     \begin{eqnarray*}
1764     \left\langle {q_\alpha ^2 } \right\rangle & = & \frac{{kT}}{{m_\alpha \omega _\alpha ^2 }} \\
1765     \left\langle {q_\alpha (t)q_\alpha (0)} \right\rangle & = & \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha t) \\
1766     \left\langle {q_\alpha (t)q_\beta (0)} \right\rangle & = &\delta _{\alpha \beta } \left\langle {q_\alpha (t)q_\alpha (0)} \right\rangle \\
1767     \left\langle {R(t)R(0)} \right\rangle & = & \sum\limits_\alpha {\sum\limits_\beta {g_\alpha g_\beta \left\langle {q_\alpha (t)q_\beta (0)} \right\rangle } } \\
1768     & = &\sum\limits_\alpha {g_\alpha ^2 \left\langle {q_\alpha ^2 (0)} \right\rangle \cos (\omega _\alpha t)} \\
1769     & = &kT\xi (t) \\
1770     \end{eqnarray*}
1771    
1772 tim 2719 Thus, we recover the \emph{second fluctuation dissipation theorem}
1773 tim 2696 \begin{equation}
1774     \xi (t) = \left\langle {R(t)R(0)} \right\rangle
1775 tim 2719 \label{introEquation:secondFluctuationDissipation}.
1776 tim 2696 \end{equation}
1777 tim 2719 In effect, it acts as a constraint on the possible ways in which one
1778     can model the random force and friction kernel.